• Aucun résultat trouvé

Inorganic carbon and nitrogen assimilation in cellular compartments of a benthic kleptoplastic foraminifer

N/A
N/A
Protected

Academic year: 2021

Partager "Inorganic carbon and nitrogen assimilation in cellular compartments of a benthic kleptoplastic foraminifer"

Copied!
13
0
0

Texte intégral

(1)

HAL Id: hal-02872797

https://hal.univ-angers.fr/hal-02872797

Submitted on 21 Dec 2020

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Distributed under a Creative Commons Attribution - NoDerivatives| 4.0 International License

Charlotte Lekieffre, Thierry Jauffrais, Emmanuelle Geslin, Bruno Jesus, Joan Bernhard, et al.. Inor- ganic carbon and nitrogen assimilation in cellular compartments of a benthic kleptoplastic foraminifer.

Scientific Reports, Nature Publishing Group, 2018, 8 (1), pp.10140. �10.1038/s41598-018-28455-1�.

�hal-02872797�

(2)

www.nature.com/scientificreports

Inorganic carbon and nitrogen assimilation in cellular

compartments of a benthic kleptoplastic foraminifer

Charlotte LeKieffre

1,2

, Thierry Jauffrais

2

, Emmanuelle Geslin

2

, Bruno Jesus

3,4

, Joan M. Bernhard

5

, Maria-Evangelia Giovani

1

& Anders Meibom

1,6

Haynesina germanica, an ubiquitous benthic foraminifer in intertidal mudflats, has the remarkable

ability to isolate, sequester, and use chloroplasts from microalgae. The photosynthetic functionality of these kleptoplasts has been demonstrated by measuring photosystem II quantum efficiency and

O2

production rates, but the precise role of the kleptoplasts in foraminiferal metabolism is poorly understood. Thus, the mechanism and dynamics of C and N assimilation and translocation from the kleptoplasts to the foraminiferal host requires study. The objective of this study was to investigate, using correlated TEM and NanoSIMS imaging, the assimilation of inorganic C and N (here ammonium, NH

4+

) in individuals of a kleptoplastic benthic foraminiferal species. H. germanica specimens were incubated for 20 h in artificial seawater enriched with H

13CO3 and 15

NH

4+

during a light/dark cycle. All specimens (n = 12) incorporated

13

C into their endoplasm stored primarily in the form of lipid droplets.

A control incubation in darkness resulted in no 13

C-uptake, strongly suggesting that photosynthesis is the process dominating inorganic C assimilation. Ammonium assimilation was observed both with and without light, with diffuse

15

N-enrichment throughout the cytoplasm and distinct

15

N-hotspots in fibrillar vesicles, electron-opaque bodies, tubulin paracrystals, bacterial associates, and, rarely and at moderate levels, in kleptoplasts. The latter observation might indicate that the kleptoplasts are involved in N assimilation. However, the higher N assimilation observed in the foraminiferal endoplasm incubated without light suggests that another cytoplasmic pathway is dominant, at least in darkness.

This study clearly shows the advantage provided by the kleptoplasts as an additional source of carbon and provides observations of ammonium uptake by the foraminiferal cell.

Kleptoplasty is defined as the process in which a cell sequesters algal chloroplasts while discarding or digesting other algal components

1

. This phenomenon is encountered in different organisms, such as sacoglossans (sea slugs)

2–4

, ciliates

5

, dinoflagellates

6–8

, and benthic foraminifera

9,10

.

Studies of benthic foraminiferal kleptoplasty have focused on shallow-water species inhabiting photic zones, especially Haynesina germanica and Elphidium spp. These studies have relied on ultrastructural observations and/or genetic analyses, which established the diatom origin of the kleptoplasts, or incubation/starvation exper- iments to define kleptoplast lifetimes and functionality once inside the foraminiferal cells

10–21

. Additionally

22

, showed that H. germanica and Elphidium williamsoni had a net uptake of inorganic carbon (H

14

CO

3

) in light, and experiments with oxygen microelectrodes demonstrated that maximal O

2

production by H. germanica depended on light intensity and light history

11,23

. A kleptoplastic strategy thus provides these organisms with both carbon

1Laboratory for Biological Geochemistry, School of Architecture, Civil and Environmental Engineering (ENAC), Ecole Polytechnique Fédérale de Lausanne (EPFL), 1015, Lausanne, Switzerland. 2UMR CNRS 6112 LPG-BIAF, Université d’Angers, 2 Boulevard Lavoisier, 49045, Angers, CEDEX 1, France. 3EA2160, Laboratoire Mer Molécules Santé, Université de Nantes, Nantes, France. 4BioISI – Biosystems & Integrative Sciences Institute, Campo Grande University of Lisboa, Faculty of Sciences, Lisboa, Portugal. 5Woods Hole Oceanographic Institution, Department of Geology & Geophysics, Woods Hole, MA, USA. 6Center for Advanced Surface Analysis, Institute of Earth Sciences, University of Lausanne, 1015, Lausanne, Switzerland. Correspondence and requests for materials should be addressed to C.L. (email: charlotte.lekieffre@epfl.ch)

Received: 18 April 2018 Accepted: 20 June 2018 Published: xx xx xxxx

OPEN

(3)

and a source of oxygen. Cesbron et al.

11

hypothesized that kleptoplasts might constitute an additional carbon source that may provide the kleptoplastic foraminifera a substantial competitive advantage, especially during periods of impoverished nutrients. However, the extent to which kleptoplasty contributes to the carbon within the foraminiferal cell storage via photosynthetic C assimilation has not been studied yet.

Foraminiferal kleptoplasts might also be involved in uptake of inorganic N. Indeed, diatoms, from which foraminifera sequester their kleptoplasts

20,24,25

, are able to assimilate ammonium through the chloroplast GS/

GOGAT (glutamate synthase and glutamine oxoglutarate aminotransferase) enzymatic pathway

26–28

. Kleptoplasty is also shown in deep-sea species

9,24,25,29

living in complete darkness and thus unable to perform photosynthe- sis

9,24

. Among these deep-sea species, Nonionella stella maintains kleptoplasts and associated functional enzy- matic machinery, including ribulose bis-phosphate carboxylase oxygenase (RuBisCO) and phosphoenol pyruvate carboxylase (PEP carboxylase), intact for months in the dark after sampling

24

. It was suggested that kleptoplasts in these deep-sea species are involved in assimilation of inorganic N

24

. Therefore, a similar role in shallow klepto- plastic foraminiferal species is possible and needs to be investigated.

To date, no studies have documented the timing and distribution of assimilation and translocation of C-photosynthates or N-compounds between kleptoplasts and the foraminiferal cell. To precisely trace the C and N assimilation within the different cell compartments in the kleptoplastic foraminiferal cell, the NanoSIMS (nanos- cale secondary ion mass spectrometry, a relatively recent ultra-high resolution isotopic imaging method

30,31

), was used in combination with transmission electron microscopy (TEM) and stable isotope labeling experiments

30,31

. This combined approach has already been successfully applied to study assimilation, storage, and transfer of C and N in several different marine organisms

32–39

, including foraminifera

40–43

. Using this integrative approach, the present study had three objectives: (1) investigate the role of kleptoplasts in C-fixation, (2) investigate the transfer and distribution of photosynthetically produced organic C within the host; and (3) investigate the potential role of kleptoplasts in foraminiferal N metabolism.

Results

H. germanica specimens were incubated for 20 h in artificial seawater enriched with 2 mM NaH

13

CO

3

and 10 μM

15

NH

4

Cl, following a light - dark cycle (Fig. 1, see details in Methods). Specimens were preserved for analysis at regular time intervals (i.e., after 4, 8, 12, and 20 hours).

TEM observations of foraminiferal cytoplasm. The cytoplasm of all specimens had well-preserved ultrastructure (Fig. 2A), as well as intact mitochondria with visible double-membranes and cristae (Fig. 2B).

Numerous small lipid droplets (diameter of ca. 500 nm), recognized by their waxy appearance, were observed in the cytoplasm (Fig. 2C), along with some larger lipid droplets ranging from 1 to 3 µm in diameter. Numerous small oval fibrillar vesicles (ca. 500 nm in length), with the fibrils arranged in parallel, and spherical to oval-shaped electron-opaque bodies (200–500 nm) were observed in the cytoplasm (Fig. 2D,E), along with occasional tubulin paracrystals identifiable due to the regular pattern of their ultrastructure revealed by high-magnification TEM imaging (Fig. 2F). In all specimens, we observed many small structures (2 to 3 µm in length) variable in shape but mainly ovoid (Fig. 2A) with the presence of numerous vacuoles within their matrix (Fig. 2G,H). Henceforth, we refer to these as “multi-vacuolar structures”.

In all observed specimens, TEM images of the endoplasm revealed well-preserved kleptoplasts with visible pyrenoids and thylakoids (Figs 2A, 3, 4). These kleptoplasts ranged in size from 2 to 10 µm in diameter. Generally, their outlines were circular to oval. They were distributed in the endoplasm with no clear pattern and often sur- rounded by an electron-lucent space between the kleptoplast membranes and the endoplasm. Some of the small lipid droplets were observed adjacent to the kleptoplast periphery. In some cases, lipid droplets were even closely associated with kleptoplast membranes (Fig. 3B).

Uptake of H

13

CO

3

within foraminiferal cells. In Experiment 1, starting at t

=

8 h,

13

C-enrichments

were detected in all specimens. The signal was concentrated in fibrillar vesicles, electron opaque bodies, and

lipids (Figs 4, 5A,B,C). In contrast, only one specimen from the first time point (i.e. at 4 h) exhibited

13

C-enriched

structures, concentrated in fibrillar vesicles and electron opaque bodies (Fig. 4). Although some lipid drop-

lets were present, they were only slightly enriched at 4 h (Figs 4 and 6). All specimens collected between 8

Figure 1. Schematic of Experiments 1 and 2, exposing H. germanica to different light conditions. Three

specimens were sampled at each indicated time point. See text for details.

(4)

www.nature.com/scientificreports/

and 20 h of incubation exhibited

13

C-enrichments in the endoplasm. The

13

C-enrichment (expressed in

δ13

C) of electron-opaque bodies significantly increased during the light phase from ca. 40‰ at 4 h to 180‰ at 8 h (p

<

0.05) and remained stable through the dark phase of the experiment, i.e. between 8 and 20 h of incuba- tion (Fig. 6B). The

13

C-enrichments of the fibrillar vesicles and the lipid droplets were relatively stable during the incubation at ca. 200 to 300‰ and 60 to 150‰, respectively (Fig. 6C,D). The cytoplasm itself was slightly more enriched after 8 h of incubation than after 4 h, with averages of ca. 100‰ and 40‰, respectively (Fig. 6A).

However, the cytoplasmic enrichment did not change statistically between 8 h and 20 h (p

>

0.05; Fig. 6A). No

13

C-enrichments were found in foraminifera incubated with H

13

CO

3

in darkness (Experiment 2; Figs 6 and 7).

Uptake of

15

NH

4+

in the foraminiferal cell. All specimens of Experiment 1 exhibited detectable

15

N-enrichments. In the cytoplasm of H. germanica,

15

N-enrichments significantly increased between 4 and 8 h Figure 2. TEM micrographs of the cytoplasm and organelles of Haynesina germanica. (A) Aspect of the cytoplasm in a chamber of the penultimate whorl. (B) Intact mitochondria with well-defined cristae and intact double-membranes. (C) Small lipid droplets. (D) Fibrillar vesicles. (E) Electron-opaque bodies, (F) Tubulin paracrystals; Inset: higher magnification revealing regular pattern of the paracrystal ultrastructural organization. (G,H) multi-vacuolar structures. Arrowheads: multi-vacuolar structures; c: chloroplast; eo:

electron-opaque bodies, fv: fibrillar vesicles, li: lipid droplets; m: mitochondria, tp: tubulin paracrystals, v-

vacuole. Scale bars: A: 2 µm; B, inset F: 200 nm; C–H: 500 nm.

(5)

(during the light phase), from ca. 250 to 550‰ (p < 0.05), and stabilized between 8 and 20 h (p > 0.05), i.e., during the dark phase (Fig. 6A). Similar to the observed

13

C-enrichments, the

15

N-signal was concentrated in electron-opaque bodies and fibrillar vesicles (Figs 4, 5A,B and 6). The

15

N-enrichment of electron-opaque bod- ies was relatively stable from 4 to 12 h (between 1800–2900‰) and then decreased to less than 1000‰ at 20 h (p < 0.05 between 4 and 20 h; Fig. 6B). At the same time, the

15

N-enrichment of the fibrillar vesicles increased throughout the incubation passing from ca. 425‰ at 4 h to 1320‰ after 20 h (p < 0.05). Some of these organelles were occasionally enriched in

15

N but not in

13

C. The tubulin paracrystals and the multi-vacuolar structures were also strongly enriched in

15

N after 8 h (Fig. 5D,E). Kleptoplasts rarely exhibited

15

N-enrichments, and if such enrichments were observed, they were always moderate to low (Fig. 5F).

In Experiment 2, after 8 h in darkness, the foraminifera had incorporated a much higher concentra- tion of

15

NH

4+

(Fig. 6) compared with Experiment 1 at any given time (p < 0.05); the cytoplasmic average

15

N-enrichment reached a value of ca. 1100‰ after 8 h of incubation in darkness. The fibrillar vesicles were also significantly more

15

N-enriched during the second experiment reaching values of ca. 2400‰ (vs. maximum of ca. 1300‰ during Experiment 1). In contrast, the electron-opaque bodies exhibited

15

N-enrichment compara- ble to those recorded during the first experiment, i.e. around ca. 2600‰. As in Experiment 1, the

15

N isotopic signal was observed most concentrated in electron-opaque bodies, fibrillar vesicles, tubulin paracrystals, in the multi-vacuolar structures, as well as in a few kleptoplasts (Figs 6 and 7).

Discussion

All specimens exhibited mitochondria with intact cristae and double-membranes indicating that they were alive at the time of fixation

42,44

. The kleptoplasts observed in our study correspond to the morphological description of

10

for H. germanica collected from the Bourgneuf Bay (as in this study) and from the Wadden Sea (Mokbaai, NL). Specimens were well preserved with undamaged thylakoids and pyrenoids. The electron-lucent space that sometimes surrounded the kleptoplasts was also previously described by

10

and ascribed to a possible fixation artefact. Indeed, TEM observations of the same species fixed using high pressure freezing and freeze substitution Figure 3. TEM micrographs of one chloroplast in Haynesina germanica cytoplasm. (A) Intact pyrenoid and thylakoids. (B) Higher magnification image showing two small lipid droplets in contact with the chloroplast membranes. The chloroplast membranes adjacent to the lipid vesicle are disrupted. li: lipid droplets, py:

pyrenoid, th: thylakoid. Scale bars: A: 2 µm; B: 500 nm.

(6)

www.nature.com/scientificreports/

instead of “classic” chemical fixation revealed kleptoplasts without electron-lucent space, with their membranes directly in contact with the surrounding foraminiferal cytoplasm

16,17

.

The paired TEM-NanoSIMS observations allowed the visualization of inorganic C uptake (H

13

CO

3

) within foraminiferal cells incubated under a light/dark cycle (Figs 4 and 6). The absence of

13

C assimilation in contin- uous darkness (Experiment 2, Figs 6 and 7) and the observed production of O

2

under light, as observed for H.

germanica in other studies

11,23

, strongly suggests that H. germanica kleptoplasts have a functional Calvin-Benson

cycle, resulting in the production and transfer of

13

C-photosynthates to the H. germanica cell. Foraminifera can

acquire C by different trophic mechanisms

45

, but they are not known to actively uptake inorganic C in the absence

of either bacterial or algal symbionts or in the presence of kleptoplasts. We found no indications of the pres-

ence of prokaryotic symbiotic photosynthetic organisms and, therefore, suggest that the observed incorpora-

tion of

13

C-bicarbonate is the result of photosynthesis occurring in the kleptoplasts. However, the absence of

Figure 4. Time-evolution of

13

C and

15

N uptake and localization within the cytoplasm of H. germanica during

Experiment 1 (light/dark incubation with H

13

CO

3

and

15

NH

4+

). Left column: TEM micrographs. Middle and

right columns: corresponding NanoSIMS

δ13

C and

δ15

N images, respectively, expressed in ‰. Arrows: fibrillar

vesicles; arrowheads: electron opaque bodies; circles and li: lipid droplets, white triangles: multi-vacuolar

structures; c: chloroplast; tp: tubulin paracrystals. Scale bars: 2 µm.

(7)

13

C-enrichment inside the kleptoplasts (Figs 4 and 5) was unexpected. This absence of

13

C-enrichment can be attributed to the fact that the

13

C-photosynthates are quickly transported away from the kleptoplasts and thus the

13

C-enrichment stay below the detection limit of the NanoSIMS. This hypothesis is supported by previous NanoSIMS studies of autotrophic

13

C-exchanges in the symbiotic association between dinoflagellates and corals, where

13

C-enrichments in dinoflagellate chloroplasts were systematically much lower than in other sub-cellular organelles

34

. Additionally, studies on H. germanica have shown that the cellular machinery necessary for chloro- plast maintenance is unlikely to be functional

23

, which could explain why, in our Experiment 1, the kleptoplasts did not accumulate

13

C within their structures. In summary, our observations show that kleptoplasts in H. ger- manica are able to assimilate inorganic C and form

13

C-photosynthates that are transferred to the host cell, but the kleptoplasts do not themselves become enriched in

13

C (above the detection limit of the NanoSIMS).

The numerous multi-vacuolar structures observed in H. germanica (Fig. 2A,G,H) are somewhat similar to the bacteria observed in another benthic species, Globocassidulina cf. G. biora

46

. The presence of numerous such Figure 5. Foraminiferal organelles enriched in

13

C and/or

15

N in Experiment 1 at different time points. Left column: TEM micrographs. Middle and right columns: corresponding NanoSIMS δ

13

C and δ

15

N images, respectively, expressed in ‰. (A) electron-opaque bodies (after 8 h of incubation), (B) fibrillar vesicles (after 12 h of incubation), (C) lipid droplets (after 8 h of incubation), (D) tubulin paracrystals (after 20 h of incubation), (E) multi-vacuolar structures (after 12 h of incubation), (F) chloroplasts (after 8 h of incubation).

Arrowheads: electron-opaque bodies; arrows: fibrillar vesicles; circles and li: lipid droplets; c: chloroplasts; white

triangles: multi-vacuolar structures; c: chloroplast; tp: tubulin paracrystals. Scale bars: 2

µ

m.

(8)

www.nature.com/scientificreports/

vacuoles within prokaryotic cells has been described and linked to different biological functions, such as buoy- ancy (gas vacuoles) in planktonic bacteria or nitrate vacuoles in filamentous sulfur bacteria

47,48

. Thus these struc- tures could potentially be interpreted as endobionts. Is it noteworthy that they were not labeled in

13

C (Fig. 5E).

Other NanoSIMS studies looking at

13

C-bicarbonate assimilation in cyanobacteria, anaerobic photosynthetic bacteria or chemotrophic bacteria have shown strong bacterial

13

C-enrichments

49–52

. Therefore, even if the multi-vacuolar structures observed in our study were bacteria, the absence of

13

C incorporation into their struc- ture suggests that they are not photosynthesizing or assimilating inorganic carbon and that thus they do not play any role in the inorganic

13

C assimilation in Haynesina germanica.

Carbon was assimilated during the light phase, transferred to the foraminiferal cell, and accumulated in spe- cific organelles: electron-opaque bodies, fibrillar vesicles, and lipid droplets. The

13

C-assimilation dynamics in Figure 6. Average

13

C (black bars) and

15

N (white bars) enrichment of (A) the cytoplasm, (B) electron-opaque bodies, (C) fibrillar vesicles and (D) lipid droplets of H. germanica (n

=

3) in Experiments 1 and 2. Error bars represent one standard deviation. Red lines indicate natural variations in

13

C (solid lines) and

15

N (dotted lines) enrichments as measured by NanoSIMS in similar areas of unlabelled control H. germanica specimens (n

=

3;

δ13

C

=

0

±

40‰, and

δ15

N

=

0

±

60‰, 3σ).

(9)

electron-opaque bodies and fibrillar vesicles should be interpreted with caution, as these organelles are poorly understood

44

. Indeed, we do not know their function(s) in the cell, how fast they are produced, their turn over and what triggers their production. In addition, we do not know if the

13

C-enriched hotspots observed within the foraminiferal endoplasm correspond only to newly formed electron-opaque bodies and fibrillar vesicles, or if

13

C-enriched material was added to pre-existing organelles; which is likely considering the high variability observed for enrichment values (Fig. 6). However, we note that none of these organelles exhibited an increase in their

13

C-enrichment during the dark phase of Experiment 1, and none of the analyzed organelles showed

13

C-enrichment during Experiment 2. This strongly indicates that there are no cytoplasmic foraminiferal path- ways for inorganic carbon assimilation; i.e. carbon is assimilated only via the photosynthetic kleptoplasts.

Lipid droplets are considered to be the main C storage form in foraminifera

44

. A similar accumulation process/

sequence has been observed in the symbiotic planktonic foraminifer Orbulina universa, where photosynthesis led to an assimilation of inorganic C (H

13

CO

3

) stored in the form of lipid droplets

43

. In kleptoplastic sea slugs (e.g., Elysia chlorotica), lipid droplets observed in the animal tissue where argued to result from the exudation of lipids from the plastids because their fatty acids had a large proportion of algal-derived eicosapentaenoic acid (20:5)

53

. However, these authors could not determine whether the plastids transferred fatty acids directly via triacylglycerols (TAGs), or as free fatty acids that may be further transformed by the host into lipid droplets. The de novo produc- tion of triacylglycerol by chloroplasts in marine algae has been demonstrated

54,55

. Furthermore, de novo fatty acid synthesis is known to occur in plant cell chloroplasts

56

, followed by a transfer in the form of free fatty acids to the cytosol

57

. Additional transfer of soluble molecules such as maltose or glucose across the chloroplast membranes through transporters also occurs in plant cells

58

. The close spatial association between kleptoplast membranes and small lipid droplets observed here (Fig. 3) may indicate a potential transfer of C via exudation of small lipid droplets from kleptoplast to the H. germanica cell, although the detailed mechanisms by which the fatty acids cross kleptoplast membranes remain unknown. Unfortunately, the distribution of soluble molecules cannot be investi- gated with NanoSIMS because the sample preparation protocol causes near complete loss of such components

41

.

Teugels et al.

59

reported that ammonium assimilation by the kleptoplastic sacoglossan Elysia viridis was significantly higher during light exposure than in darkness. This is consistent with the glutamine oxoglutarate aminotransferase (GOGAT) enzyme pathway that requires electron donors (e.g., reduced ferredoxin) formed during photosynthetic electron transport

60

. Furthermore, the glutamate synthase (GS) metabolic reaction is ATP-dependent, and gene expression of some key enzymes (GS and GOGAT) is light regulated

60

. In corals, sym- biotic dinoflagellate GS/GOGAT enzymes are thought to be the main ammonium assimilation pathway

35,38,61,62

. Furthermore, cnidarian cells are also known to produce cytosol glutamate dehydrogenase (GDH)

62–64

, which has a dual function: 1) it converts glutamate to

α

-ketoglutarate and ammonium that is subsequently assimi- lated into chloroplasts via the GOGAT pathway

59

; 2) it catalyzes the opposite reaction, i.e. the amination of the

α

-ketoglutarate to produce the amino acid glutamate

65

.

In our study, the observation of

15

N-labeled kleptoplasts in H. germanica incubated both in light and in darkness seems consistent with a GS/GOGAT kleptoplastic pathway for ammonium assimilation (Fig. 5H). However, the uptake of

15

N-ammonium was higher after 8 h of incubation in total darkness (Experiment 2) than in Experiment 1 (light-dark cycle) (Fig. 6). This higher uptake in darkness compared to light is inconsistent with the light regu- lation of the GS/GOGAT enzymatic machinery

60

. Ammonium incorporation might, thus, also occur by another N-assimilation pathway in foraminifera, for example, through the glutamine dehydrogenase (GDH) pathway.

Figure 7.

13

C and

15

N uptake and localization within the cytoplasm of H. germanica during Experiment 2

(continuous dark incubation with H

13

CO

3

and

15

NH

4+

). Left column: TEM micrographs. Middle and right

columns: corresponding NanoSIMS

δ13

C and

δ15

N images, respectively, expressed in ‰. Arrows: fibrillar

vesicles; arrowheads: electron opaque dense bodies; circles: lipid droplets, white triangles: multi-vacuolar

structures; c: chloroplast, tp: tubulin paracrystals. Scale bars: 2 µm.

(10)

www.nature.com/scientificreports/

The same organelles (fibrillar vesicles, electron-opaque bodies and multi-vacuolar structures) were found to be

15

N-enriched in both Experiments 1 and 2. As previously noted for

13

C-assimilation, the

15

N-assimilation dynamics in electron-opaque bodies and fibrillar vesicles is difficult to interpret, due to the lack of knowledge about the function of these organelles

44

. However, the electron-opaque bodies and the fibrillar vesicles seem to have different patterns for

15

N assimilation. While the electron-opaque bodies incorporated large amounts of

15

N even after only 4 hours and then gradually lost this

15

N-enrichment over time, the fibrillar vesicles assimilated

15

N throughout the 20 h incubation, independent of light condition. It is noteworthy that the

15

N-assimilation pattern clearly differs from the

13

C-assimilation dynamics in both organelle types.

Finally, ammonium is known to be a suitable N source for many marine prokaryotes

66–68

. Thus, if the multi-vacuolar structures, abundant in all H. germanica specimens (Fig. 2A,G,H) are endosymbionts (see above) they would be expected to incorporate

15

NH

4+

, as is indeed observed (Fig. 5G). They could, thus, constitute another putative nitrogen assimilation pathway for the benthic foraminifer H. germanica. However, we cannot conclude further in this study about the symbiotic nature of these putative prokaryotes in H. germanica, nor about their role in foraminiferal N metabolism.

A comparative study of organic C (algae) uptake through feeding between the two dominant foraminiferal species inhabiting mudflats, the akleptoplastic Ammonia sp. and kleptoplastic H. germanica, showed a higher uptake rate by the former

69

. Our results highlight that H. germanica can fix inorganic carbon. Therefore, unlike Ammonia sp., H. germanica does not rely solely on heterotrophy to meet its C requirements. The mixotrophic feeding strategy of H. germanica might give a competitive advantage and allow it to become the dominant foraminifera in mudflat environments

70–72

. In addition, whether H. germanica assimilates nitrogen through the kleptoplasts, potential endosymbionts, and/or by another pathway specific to foraminifera, our observations demonstrate that it is also capable of using inorganic N as a nutrient source. Further investigation is required to quantify this uptake and elucidate the role of this benthic foraminifera in the N cycle, especially since H. german- ica thrives in coastal ecosystems that are subject to increasing eutrophication

73,74

.

Conclusion

Our study demonstrates inorganic C assimilation in H. germanica, most likely via the kleptoplasts. The absence of

13

C assimilation in darkness combined with previous studies documenting O

2

production in light strongly suggest that photosynthesis is the process dominating inorganic C-assimilation in this species. Subsequently, photosynthates are transferred to the foraminiferal cell and utilized for its metabolism. Thus, these observations clearly show the role played by the kleptoplasts in H. germanica carbon metabolism, providing the foraminiferal cell with an additional autotrophic source of C. The observation of small lipid droplets attached to the klepto- plast membranes may suggest a transfer of C from the kleptoplasts to the foraminiferal cell in the form of lipids.

However, the detailed mechanism(s) involved in this C transfer remains unknown. The kleptoplasts may also pro- vide additional N sources to foraminiferal metabolic pathways via GS-GOGAT enzymes. However, ammonium assimilation was more efficient in darkness, requiring the existence of other N-assimilation pathways.

Material and Methods

Experiment 1: light/dark cycle incubation with H

13

CO

3

and

15

NH

4+

. Living foraminifera were collected on April 9, 2015, at low tide on the intertidal mudflat of the Bourgneuf Bay (France, 47°00

59.4

N 2°01

29.8

W). The top centimeter of the sediment was sampled, sieved over a mesh of 150

μ

m with in situ seawa- ter and the

>

150

μ

m fraction was immediately transported in the dark over

3 hours to the laboratory in Nantes.

In the laboratory, healthy living individuals of H. germanica were selected under a binocular microscope based on their cytoplasm color (i.e. yellow-brownish material spread through all the chambers of the specimen, except the youngest chamber). The selected specimens were placed into 5 Petri dishes (5 specimens per Petri dish) filled with artificial seawater (ASW, Red Sea Salt, salinity

=

35, pH

=

8.0). Four of the Petri dishes contained ASW enriched with 2 mM NaH

13

CO

3

and 10

μ

M

15

NH

4

Cl. The fifth Petri dish contained isotopically normal artificial seawater of the same chemical composition: specimens from this dish were fixed at T0 (beginning of the experiment) and served as controls for NanoSIMS analysis (see below). All other Petri dishes were placed in an incubator (Fytoscope FS130, temperature: 18 °C, light intensity: 90

μ

mol m

−2

s

−1

). After 8 h of light exposure they were transferred to dark conditions for another 12 h. Except for the control specimens, the foraminifera remained in the spiked ASW throughout the experiment. At each time point, i.e., after 4, 8, 12, and 20 hours, one Petri dish was removed from the incubator (Fig. 1) and the 5 specimens contained in this Petri dish were immediately chemically fixed.

Experiment 2: Incubation in continuous darkness with H

13

CO

3

and

15

NH

4+

. H. germanica speci- mens were collected on May 16, 2015, at low tide on the intertidal mudflat of the Bourgneuf Bay (France) follow- ing the same procedure as described above. Five living specimens were selected and placed in a Petri dish with artificial seawater (Red Sea Salt, salinity

=

35; pH

=

8.0) enriched with 2 mM NaH

13

CO

3

and 10

μ

M of

15

NH

4

Cl.

They were incubated in continuous darkness for 8 h (Fig. 1) and immediately chemically fixed at the end of this incubation. Control samples, which were incubated in normal seawater, where fixed at the beginning of the experiment (T0; Fig. 1).

Preparation for TEM-NanoSIMS analysis. The specimens were chemically fixed following the pro-

tocol described

44

). Briefly, foraminifera were fixed immediately after removal from the incubator, with a mix

of 4% glutaraldehyde and 2% paraformaldehyde diluted in 0.1 M cacodylate buffer, 0.4 M sucrose, and 0.1 M

NaCl (pH

=

7.4) at room temperature for 24 h. They were then stored at 4 °C until further processing. Further

chemical processing and transmission electron microscope (TEM) imaging of the foraminifera were performed

at the Electron Microscopy Facility of the University of Lausanne (Switzerland). After rinsing, specimens were

(11)

Foraminiferal sections were imaged with the NanoSIMS ion microprobe with a 16 keV primary ion beam of Cs

+

focused to a beam spot of ca. 100–150 nm. The secondary molecular ions

12

C

2

,

13

C

12

C

,

12

C

14

N

and

12

C

15

N

were collected simultaneously in electron multiplier detectors at a mass-resolution of ca. 10000, enough to resolve potential interferences in the mass spectrum

34,35

. Isotopic image dimensions ranged from 15

×

15 µm to 30

×

30 µm with 256

×

256 pixel resolution. For each image, 6 layers were acquired, drift corrected and super- imposed using the software L’IMAGE (developed by Dr. Larry Nittler, Carnegie Institution of Washington DC, USA). Quantified

13

C/

12

C and

15

N/

14

N ratios were obtained as follows:

C (‰) (( C

mes

/ C

nat

) 1) 10

13 3

δ = − ×

δ15

N (‰)

=

(( N

mes

/ N

nat

)

1)

×

10

3

where C

mes

is the measured

12

C

13

C

/

12

C

2

ratio of the sample and C

nat

is the average

12

C

13

C

/

12

C

2

ratio measured in unlabeled samples (control). Similarly, N

mes

is the measured

12

C

15

N

/

12

C

14

N

ratio of the sample and N

nat

is the average

12

C

15

N

/

12

C

14

N

ratio measured in unlabeled samples. The software Look@NanoSIMS

75

was used to determine the isotopic enrichment of specific organelles that were identified morphologically from TEM images.

Regions of interest (ROIs) to quantify the average isotopic enrichment of the organelles were defined from the TEM images previously aligned with the NanoSIMS images (based on the

12

C

14

N image). For the average iso- topic enrichment of the cytoplasm, three circles of ca. 2 μm in diameter were drawn per image, avoiding highly

15

N-enriched organelles. For each specimen, between one and three NanoSIMS images were analyzed.

Statistical analysis. For each time point, three specimens were analyzed.

δ13

C and

δ15

N values for main orga- nelles and cytoplasm were obtained by calculating the average of ROIs within each specimen, and then calculating the average of the three specimens for each time point. The error bars provided are thus standard deviations rep- resenting inter-specimen variability. However, for the statistical analysis, a linear mixed-effects (LME) model was constructed using all the ROIs of the three specimens for each time point (taking into account pseudo-replication effects, i.e. regrouping ROIs from three different specimens into one category), followed by a Tukey multiple comparison test. The results of the Tukey multiple comparisons tests are given in the Supplementary Data S1. The statistical analyses were performed with Rstudio software using a significance level set at α

=

0.05.

Data Availability. The datasets generated and/or analyzed during the current study are available in the PANGAEA Repository (https://doi.org/10.1594/PANGAEA.891407 data will be published upon acceptance of the manuscript).

References

1. Clark, K. B., Jensen, K. R. & Stirts, H. M. Survey for functional kleptoplasty among west Atlantic Ascoglossa (=Sacoglossa) (Mollusca: Opisthobranchia). The Veliger 33, 339–345 (1990).

2. Pelletreau, K. N. et al. Sea slug kleptoplasty and plastid maintenance in a metazoan. Plant Physiol. 155, 1561–1565 (2011).

3. De Vries, J., Rauch, C., Gregor, C. & Gould, S. B. A sea slug’s guide to plastid symbiosis. Acta Soc. Bot. Pol. 83, 415 (2014).

4. Serôdio, J., Cruz, S., Cartaxana, P. & Calado, R. Photophysiology of kleptoplasts: photosynthetic use of light by chloroplasts living in animal cells. Phil Trans R Soc B 369, 20130242 (2014).

5. Moeller, H. V. & Johnson, M. D. Preferential Plastid Retention by the Acquired Phototroph Mesodinium chamaeleon. J. Eukaryot.

Microbiol. 65, 148–158

6. Kim, M., Kim, S., Yih, W. & Park, M. G. The marine dinoflagellate genus Dinophysis can retain plastids of multiple algal origins at the same time. Harmful Algae 13, 105–111 (2012).

7. Nagai, S., Nitshitani, G., Tomaru, Y., Sakiyama, S. & Kamiyama, T. Predation by the toxic dinoflagellate Dinophys fortii on the ciliate Myrionecta rubra and observation of sequestration of ciliate chloroplasts. J. Phycol. 44, 909–922 (2008).

8. Nishitani, G. et al. Multiple plastids collected by the dinoflagellate Dinophysis mitra through kleptoplastidy. Appl. Environ. Microbiol.

78, 813–821 (2012).

9. Bernhard, J. M. & Bowser, S. S. Benthic foraminifera of dysoxic sediments: chloroplast sequestration and functional morphology.

Earth-Sci. Rev. 46, 149–165 (1999).

10. Jauffrais, T. et al. Ultrastructure and distribution of kleptoplasts in benthic foraminifera from shallow-water (photic) habitats. Mar.

Micropaleontol. 138, 46–62 (2018).

11. Cesbron, F. et al. Sequestered chloroplasts in the benthic foraminifer Haynesina germanica: cellular organization, oxygen fluxes and potential ecological implications. J. Foraminifer. Res. 47, 268–278 (2017).

12. Cevasco, M. E., Lechliter, S. M., Mosier, A. E. & Perez, J. Initial observations of kleptoplasty in the foraminifera of coastal South Carolina. Southeast. Nat. 14, 361–372 (2015).

13. Correia, M. J. & Lee, J. J. Fine structure of the plastids retained by the foraminifer Elphidium excavatum (Terquem). Symbiosis 32, 15–26 (2002).

(12)

www.nature.com/scientificreports/

14. Correia, M. J. & Lee, J. J. How long do the plastids retained by Elphidium excavatum (Terquem) last in their host? Symbiosis 32, 27–37 (2002).

15. Correia, M. J. & Lee, J. J. Chloroplast retention by Elphidium excavatum (Terquem). Is it a selective process? Symbiosis 29, 343–355 (2000).

16. Goldstein, S. T., Bernhard, J. M. & Richardson, E. A. Chloroplast sequestration in the foraminifer Haynesina germanica: Application of high pressure freezing and freeze substitution. Microsc. Microanal. 10, 1458–1459 (2004).

17. Goldstein, S. T. & Richardson, E. A. Fine structure of the foraminifer Haynesina germanica (Ehrenberg) and its sequestered chloroplasts. Mar. Micropaleontol. 138, 63–71 (2018).

18. Jauffrais, T., Jesus, B., Méléder, V. & Geslin, E. Functional xanthophyll cycle and pigment content of a kleptoplastic benthic foraminifer: Haynesina germanica. PLOS ONE 12, e0172678 (2017).

19. Lee, J. J. & Lanners, E. The retention of chloroplasts by the foraminifer Elphidium crispum. Symbiosis 5, 45–59 (1988).

20. Pillet, L., de Vargas, C. & Pawlowski, J. Molecular identification of sequestered diatom chloroplasts and kleptoplastidy in foraminifera. Protist 162, 394–404 (2011).

21. Pillet, L. & Pawlowski, J. Transcriptome analysis of foraminiferan Elphidium margaritaceum questions the role of gene transfer in kleptoplastidy. Mol. Biol. Evol. 30, 66–69 (2013).

22. Lopez, E. Algal chloroplasts in the protoplasm of three species of benthic foraminifera: taxonomic affinity, viability and persistence.

Mar. Biol. 53, 201–211 (1979).

23. Jauffrais, T. et al. Effect of light on photosynthetic efficiency of sequestered chloroplasts in intertidal benthic foraminifera (Haynesina germanica and Ammonia tepida). Biogeosciences 13, 2715–2726 (2016).

24. Grzymski, J., Schofield, O. M., Falkowski, P. G. & Bernhard, J. M. The function of plastids in the deep-sea benthic foraminifer.

Nonionella stella. Limnol. Oceanogr. 47, 1569–1580 (2002).

25. Tsuchiya, M. et al. Cytologic and Genetic Characteristics of Endobiotic Bacteria and Kleptoplasts of Virgulinella fragilis (Foraminifera). J. Eukaryot. Microbiol. n/a-n/a. https://doi.org/10.1111/jeu.12200 (2015).

26. Syrett, P. J. Nitrogen metabolism of microalgae. Can. Bull. Fish. Aquat. Sci. (1981).

27. Zehr, J. P., Falkowski, P. G., Fowler, J. & Capone, D. G. Coupling between ammonium uptake and incorporation in a marine diatom:

experiments with the short-lived radioisotope 13N. Limnol. Oceanogr. 33, 518–527 (1988).

28. Zehr, J. P. & Falkowski, P. G. Pathway of Ammonium Assimilation in a Marine Diatom Determined with the Radiotracer 13n1. J.

Phycol. 24, 588–591 (1988).

29. Cedhagen, T. Retention of chloroplasts and bathymetric distribution in the Sublittoral Foraminiferan Nonionellina labradorica.

Ophelia 33, 17–30 (1991).

30. Hoppe, P., Cohen, S. & Meibom, A. NanoSIMS: Technical aspects and applications in cosmochemistry and biological geochemistry.

Geostand. Geoanalytical Res. 37, 111–154 (2013).

31. Nuñez, J., Renslow, R., Cliff, J. B. & Anderton, C. R. NanoSIMS for biological applications: Current practices and analyses.

Biointerphases 13, 03B301 (2018).

32. Ceh, J. et al. Nutrient cycling in early coral life stages: Pocillopora damicornis larvae provide their algal symbiont (Symbiodinium) with nitrogen acquired from bacterial associates. Ecol. Evol. 3, 2393–2400 (2013).

33. Clode, P. L., Stern, R. A. & Marshall, A. T. Subcellular imaging of isotopically labeled carbon compounds in a biological sample by ion microprobe (NanoSIMS). Microsc. Res. Tech. 70, 220–229 (2007).

34. Kopp, C. et al. Subcellular investigation of photosynthesis-driven carbon assimilation in the symbiotic reef coral Pocillopora damicornis. mBio 6, e02299–14 (2015).

35. Kopp, C. et al. Highly dynamic cellular-level response of symbiotic coral to a sudden increase in environmental nitrogen. mBio 4, e00052–13 (2013).

36. Krueger, T. et al. Common reef-building coral in the Northern Red Sea resistant to elevated temperature and acidification. Open Sci.

4, 170038 (2017).

37. Krupke, A. et al. The effect of nutrients on carbon and nitrogen fixation by the UCYN-A–haptophyte symbiosis. ISME J. 9, 1635–1647 (2015).

38. Pernice, M. et al. A single-cell view of ammonium assimilation in coral–dinoflagellate symbiosis. ISME J. 6, 1314–1324 (2012).

39. Raina, J.-B. et al. Subcellular tracking reveals the location of dimethylsulfoniopropionate in microalgae and visualises its uptake by marine bacteria. eLife 6, e23008 (2017).

40. LeKieffre, C. et al. Surviving anoxia in marine sediments: The metabolic response of ubiquitous benthic foraminifera (Ammonia tepida). PLoS ONE 12, e0177604 (2017).

41. Nomaki, H. et al. Innovative TEM-coupled approaches to study foraminiferal cells. Mar. Micropaleontol. 138, 90–104 (2018).

42. Nomaki, H. et al. Intracellular isotope localization in Ammonia sp. (Foraminifera) of oxygen-depleted environments: Results of nitrate and sulfate labeling experiments. Front. Microbiol. 163, https://doi.org/10.3389/fmicb.2016.00163 (2016).

43. LeKieffre, C. et al. Assimilation, translocation, and utilization of carbon between photosynthetic symbiotic dinoflagellates and their planktic foraminifera host. Mar. Biol. 165 (2018).

44. LeKieffre, C. et al. An overview of cellular ultrastructure in benthic foraminifera: New observations of rotalid species in the context of existing literature. Mar. Micropaleontol. 138, 12–32 (2018).

45. Goldstein, S. T. Foraminifera: a biological overview. in Modern foraminifera (ed. Sen Gupta, B. K.) 37–55 (Springer-Verlag New- York, 1999).

46. Bernhard, J. M. Experimental and field evidence of Antarctic foraminiferal tolerance to anoxia and hydrogen sulfide. Mar.

Micropaleontol. 20, 203–213 (1993).

47. Jørgensen, B. B. & Gallardo, V. A. Thioploca spp.: filamentous sulfur bacteria with nitrate vacuoles. FEMS Microbiol. Ecol. 28, 301–313 (1999).

48. Walsby, A. E. Structure and function of gas vacuoles. Bacteriol. Rev. 36, 1–32 (1972).

49. Behrens, S. et al. Linking microbial phylogeny to metabolic activity at the single-cell level by using enhanced element labeling- catalyzed reporter deposition fluorescence in situ hybridization (EL-FISH) and NanoSIMS. Appl. Environ. Microbiol. 74, 3143–3150 (2008).

50. Finzi-Hart, J. A. et al. Fixation and fate of C and N in the cyanobacterium Trichodesmium using nanometer-scale secondary ion mass spectrometry. Proc. Natl. Acad. Sci. 106, 6345–6350 (2009).

51. Musat, N. et al. A single-cell view on the ecophysiology of anaerobic phototrophic bacteria. Proc. Natl. Acad. Sci. 105, 17861–17866 (2008).

52. Volland, J.-M. et al. NanoSIMS and tissue autoradiography reveal symbiont carbon fixation and organic carbon transfer to giant ciliate host. ISME J. 12, 714–727 (2018).

53. Pelletreau, K. N., Weber, A. P. M., Weber, K. L. & Rumpho, M. E. Lipid accumulation during the establishment of kleptoplasty in Elysia chlorotica. PLOS ONE 9, e97477 (2014).

54. Fan, J., Andre, C. & Xu, C. A chloroplast pathway for the de novo biosynthesis of triacylglycerol in Chlamydomonas reinhardtii. FEBS Lett. 585, 1985–1991 (2011).

55. Merchant, S. S., Kropat, J., Liu, B., Shaw, J. & Warakanont, J. TAG, You’re it! Chlamydomonas as a reference organism for understanding algal triacylglycerol accumulation. Curr. Opin. Biotechnol. 23, 352–363 (2012).

(13)

64. Yellowlees, D., Rees, T. A. V. & Fitt, W. K. Effect of ammonium-supplemented seawater on glutamine synthetase and glutamate dehydrogenase activities in host tissue and zooxanthellae of Pocillopora damicornis and on ammonium uptake rates of the zooxanthellae. Pac. Sci. 48, 291–295 (1994).

65. Srivastava, H. S. & Singh, R. P. Role and regulation of L-glutamate dehydrogenase activity in higher plants. Phytochemistry 26, 597–610 (1987).

66. Tupas, L. & Koike, I. Simultaneous uptake and regeneration of ammonium by mixed assemblages of heterotrophic marine bacteria.

Mar Ecol Prog Ser 70, 273–282 (1991).

67. Wheeler, P. A. & Kirchman, D. L. Utilization of inorganic and organic nitrogen by bacteria in marine systems. Limnol. Oceanogr. 31, 998–1009 (1986).

68. Zehr, J. P. & Ward, B. B. Nitrogen cycling in the ocean: New perspectives on processes and paradigms. Appl. Environ. Microbiol. 68, 1015–1024 (2002).

69. Wukovits, J., Enge, A. J., Wanek, W., Watzka, M. & Heinz, P. Effect of increased temperature on carbon and nitrogen uptake of two intertidal foraminifera Ammonia tepida and Haynesina germanica. Biogeosciences Discuss. 1–25, https://doi.org/10.5194/bg-2016- 509 (2016).

70. Cesbron, F. et al. Vertical distribution and respiration rates of benthic foraminifera: contribution to aerobic remineralization in intertidal mudflats covered by Zostera noltei meadows. Estuar. Coast. Shelf Sci. 179, 23–38 (2016).

71. Debenay, J.-P., Bicchi, E. & Goubert, E. & Armynot du Châtelet, E. Spatio-temporal distribution of benthic foraminifera in relation to estuarine dynamics (Vie estuary, Vendée, W France). Estuar. Coast. Shelf Sci. 67, 181–197 (2006).

72. Mojtahid, M. et al. Spatial distribution of living (Rose Bengal stained) benthic foraminifera in the Loire estuary (western France). J.

Sea Res. 118, 1–16 (2016).

73. Diaz, R. J. & Rosenberg, R. Spreading dead zones and consequences for marine ecosystems. Science 321, 926–929 (2008).

74. Zhang, J. et al. Natural and human-induced hypoxia and consequences for coastal areas: synthesis and future development.

Biogeosciences 7, 1443–1467 (2010).

75. Polerecky, L. et al. Look@NanoSIMS – a tool for the analysis of nanoSIMS data in environmental microbiology. Environ. Microbiol.

14, 1009–1023 (2012).

Acknowledgements

The electron microscopy platform at the University of Lausanne (Switzerland) is thanked for technical assistance.

This work was supported by the Swiss National Science Foundation (grant no. 200021_149333) and was part of the CNRS EC2CO-Lefe project ForChlo. It was also supported by the Region Pays de la Loire (Post-doc position of TJ, on FRESCO project) as well as the WHOI Robert W. Morse Chair for Excellence in Oceanography and The Investment in Science Fund at WHOI.

Author Contributions

C.L., M.E.G. and E.G. collected the samples from the field. C.L., E.G. and A.M. designed the experiment. C.L. and M.E.G. analysed the samples. C.L., T.J., E.G., B.J., J.M.B. and A.M. interpreted the data and discuss the results.

C.L. performed the statistical analysis, wrote the manuscript and prepared the figures. All authors reviewed the manuscript.

Additional Information

Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-018-28455-1.

Competing Interests: The authors declare no competing interests.

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre- ative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not per- mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

© The Author(s) 2018

Références

Documents relatifs

Black arrowheads: endoplasmic reticulum, arrows: nuclear envelope, g: Golgi apparatus, m: mitochondria, mvb: multivesicular bodies, n: nucleus, nu: nucleolus, li: lipid droplets,

En résumé, l’amélioration du taux de conformité globale totale aux recommandations du guide de pratique PAC n’a pas entraîné d’augmentation des

Speci fic intestinal KO mice for ATGL had increased intracellular LDs in the intestinal epithelium but unchanged TG absorption from intragastric administration of triolein,

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

In this thesis, I present new evidence on upper mantle seismic anisotropy around La Réunion hotspot – that is proposed to be fed by a deep-rooted mantle plume – and its

interprotofilament shear [33,34] and tubulin conforma- tional switching [35]) have been recently used to explain the surprising length dependent stiffness in clamped MT

Let us consider the two tasks of the algorithm. In the case that the local virtual clock triggers task Time Monitor to send messages, the host processor will need to send

Deposit of the phospholipid mixture diluted in chloroform on the conductive face of the ITO coated glass plate (step #6); Lipid mixture is dried under vacuum during 30 min (step